Skip to main content

Full text of "Spontaneous Symmetry Breaking in Quantum Systems. A review for Scholarpedia"

See other formats


J2' 

9i 



Spontaneous Symmetry Breaking in 
Quantum Systems. A review for 
Scholarpedia 

F. Strocchi 
INFN, Sezione di Pisa, Pisa Italy 

Abstract 

The mechanism of spontaneous symmetry breaking in quan- 
tum systems is briefly reviewed, rectifying part of the standard 
wisdom on logical and mathematical grounds. The crucial role 
^ r*| of the localization properties of the time evolution for the con- 

Q-f elusion of the Goldstone theorem is emphasized. 



1. Infinitely extended quantum systems 

In the development of theoretical physics, the standard way of de- 
scribing a broken symmetry has been that of introducing an explicit 
non-symmetric term in the equations of motion. A real revolution 
occurred with the realization of a much more economical and power- 
ful mechanism, called spontaneous symmetry breaking, by which 
symmetry breaking may be realized even if the equations of motion are 
symmetric (Heisenberg 1928, Nambu 1960, Goldstone 1961). 

The mechanism of spontaneous symmetry breaking for classical sys- 
tems has a counterpart for quantum systems. For such a transcription, 
in analogy with the classical crucial role is played by the infinite 

extension of the system and locality, so that each ground state defines 
a physically disjoint realization of the system (i.e. a disjoint physical 
world or phase). For this purpose, it is useful to recall that infinitely 
extended quantum systems are conveniently described by algebras A of 
local operators (Haag and Kastler 1964). 



1 



2 



For the sake of concreteness, for infinitely extended non-relativistic 
systems which allow a description in terms of canonical variables, one 
can take the algebra A of localized canonical variables generated by 
ip(f) = J d 3 xip(x)f(x.), where is the canonical field "localized" in 
x, 

[ V(x), V(y) ] = o, [V(x), V*(y) ] = *(x - y), 

and / G (^(R 3 ), of compact support (i.e. G £>(R 3 )) or of fast decrease 
(i.e. G S(R 3 )). ^(x) and ^*(x) have the meaning of destruction and 
creation operators and in the Fock representation they respectively de- 
stroy and create elementary excitations localized at x. For relativistic 
quantum fields, one can take the local field algebra J 7 generated by the 
local fields tp(f) = J d 4 xtp(x) f(x), f G 5(R 4 ). In both cases, space 
and time translations are assumed to be well defined transformations 
on the local algebras. 

In the case of a finite number 2N of degrees of freedom, described by 
the canonical variables qk,Pk, k — 1, ...N, (hereafter referred to as the 
finite dimensional case) , under general regularity conditions the Stone- 
von Neumann theorem (Stone 1930, von-Neumann 1931) states that the 
Schroedinger representation is the unique irreducible representation of 
the algebra A c of canonical variables, up to unitary equivalence (see 
below) . 

On the other hand, the local algebras A which describe infinitely 
extended systems have many inequivalent representations and the first 
step is the selection of those which are physically relevant. We re- 
call that a representation it (A) of A is a homomorphism of A (i.e. a 
mapping which preserves all the algebraic relations) into an algebra 
of operators in a Hilbert space V, w , (with the technical condition of a 
common invariant dense domain). 

Strong physical considerations motivate the following general con- 
ditions (in the case of zero temperature considered below). Such con- 
ditions apply to both non-relativistic and relativistic systems; for the 
latter ones, further conditions are needed corresponding to relativis- 
tic invariance and microscopic causality or (relativistic) locality (see 
below) . 

In the following, when we are dealing with a given representation 
n, in order to simplify the notation we shall often denote by A the 
representative n(A) of the element A. 



3 



I. (Existence of energy and momentum) In the representation ir 
of A in terms of operators in a Hilbert space Tin, the space and time 
translations, a x , a t , are described by strongly continuous groups of 
unitary operators C/(x), U(t), x G R s , t G R: 

a x (A) = f/(x)Af/(x)-\ a t (A) = l/(t)Al7(t)~\ VA G A 

The strong continuity is equivalent to the existence of the genera- 
tors, namely of the momentum P and of the energy H . 

II. (Stability or spectral condition) The energy spectrum a(H) is 
bounded from below. The relativistic invariant form of such a condition 
is a{H) > 0, H 2 - P 2 > 0. 

Clearly, the failure of the spectral condition implies that the system 
may undergo a decay into states of lower and lower energy, i.e. a 
collapse. 

III. (Ground state) Inf a(H) is a proper (non-degenerate) eigenvalue 
of H, the corresponding eigenvector \l/ is called the ground state and 
it is a cyclic vector for n(A), i.e. n(A) is the common dense domain. 
\l/o is also assumed to be invariant under space translations (or at least 
under a subgroup of them, but for simplicity, in the following, we shall 
restrict our discussion to the case of invariance under the full translation 
group) 

IV. (Local structure) Asymptotic abelianess: for any two elements 

A,B e n (A) 

weak— lim [a x (A),B] — 0, 

|x|— s>oo 

(the weak limit means the limit of the matrix elements for any pair of 
vectors belonging to T-L n ) 

If A and B describe quantities which can be measured, briefly if 
they are observable elements, asymptotic abelianess means that the 
measurement of B becomes compatible with the measurement of A (in 
the quantum mechanical sense), when the latter is translated at infinite 
(space) distance. 

Equivalently, the local structure may be codified by the cluster prop- 
erty: 

lim [{Aa x (B)) - (A) (B)} = 0, VA, B G n(A), 

|x|— >oo 

where the bracket ( ) denotes the ground state expectation, i.e. (A) = 
(*o,^*o)- 



4 



The physical meaning of the cluster property is that a decorrelation 
occurs for infinitely separated elements of the local algebra, as it must 
be for an acceptable physical interpretation. A very important result is 
that the validity of the cluster property is equivalent to the uniqueness 
of the translationally invariant state in H^. For brevity, a representa- 
tion with a unique translationally invariant ground state shall be called 
a pure phase. 

In conclusion, an infinitely extended quantum system is conveniently 
described by the local algebra A and by the set of its physically relevant 
representations; we denote by E the set of the corresponding states. 

Inequivalent representations of A describe disjoint realizations of 
the system and the possibility arises of the mechanism of symmetry 
breaking, as in the classical case. 

2. Symmetries of a quantum system and their spontaneous breaking 

In the case of finite dimensional quantum systems, described by the 
Schroedinger representation, as clarified by Wigner (Wigner 1959), a 
symmetry g is an invertible mapping of the states, i.e. of the rays, of 
the Schroedinger Hilbert space %s an d, as a consequence of Wigner 
theorem (Wigner 1959), it can be implemented by a unitary operator 
U(g) in Hs- Hence, g defines an algebraic mapping a g of the algebra 
of canonical variables A c (since A c is faithfully represented in Us)'- 

a g {A) = U{g)AU{g)-\ \/A e A c , 

which preserves all the algebraic relations (technically a g defines an 
automorphism of A c ), i.e. an algebraic symmetry. 

The relevant point is that, in the case of infinitely extended sys- 
tems there may be automorphisms a g of the local algebra A which are 
not described by unitary operators in one representation 7r of A. This 
means that g exists as a symmetry at the algebraic level, but it is not 
a symmetry of the realization of the system provided by the represen- 
tation 7r. In this case, the symmetry g is broken in H^. As in the 
classical case, this very deep mechanism cannot be reduced to the state- 
ment that "spontaneous symmetry breaking occurs in a situation where, 
given a symmetry of the equation of motion, solutions exists which are 
not invariant under the action of this symmetry without any explicit 
asymmetric input" (Stanford Encyclopedia of Philosophy 2008). The 



5 



symmetry is unbroken in the realization of the system described by 
a representation ir if it is implemented by unitary operators in H n . 

In the following, we consider the case of internal symmetries, i.e. 
symmetries which commute with space and time translations. For them 
one has the following criterion of symmetry breaking, which is at the 
basis of the standard discussion. 

Theorem 0.1 Given an internal symmetry (3, i.e. an automorphism 
of A, which commutes with space and time translations, a necessary and 
sufficient condition for (3 being broken in a representation ir, satisfying 
conditions I-IV, is that there exists A G k(A) such that {(3(A)) ^ (A). 
In this case (A) is called a symmetry breaking order parameter. 

It is worthwhile to remark that the validity of the cluster property, 
i.e. the uniqueness of the translationally invariant ground state in % n , 
is crucial for the effectiveness of the criterium. In fact, the cluster 
property is not satisfied in a representation defined by a mixed state 
and the invariance of the expectations of a mixed ground state does 
not exclude that the symmetry is broken in each pure phase in which 
the representation decomposes. 

3. Goldstone theorem without assuming relativistic invariance 

A very important consequence of the breaking of a continuous one- 
parameter Lie group of symmetries (3 X , A G R, in a representation % n , 
is the Goldstone theorem (Goldstone 1961, see also Nambu 1960). 
The theorem provides exact information on the energy momentum spec- 
trum of the intermediate states which saturate the two point function 
(j/j,{x) A), where A defines the symmetry breaking order parameter and 
jfi is the conserved current whose existence follows from the symmetry 
of the dynamics (typically from the invariance of the Lagrangian under 
(3 X ), hereafter referred to as the conserved Noether current. 

The formulation and proof in the non-relativistic case has been the 
subject of debate, in our opinion because it has not been clearly realized 
that the crucial hypothesis is a sufficient locality of the dynamics. Since 
there are interesting physical systems which exhibit the breaking of 
a continuous symmetry with an apparent evasion of the conclusions 
of the Goldstone theorem (like superconductivity, Coulomb systems 
and plasmons, Higgs mechanism etc.), it is worthwhile to examine the 
hypotheses with special care, even at the risk of looking pedantic. 



6 



The original argument by Goldstone (Goldstone 1961) relied on a 
semi-classical approximation based on a mean field ansatz. The sub- 
stantial improvement leading to a proof of the theorem by Goldstone, 
Salam and Weinberg (GSW) (Goldstone, Salam, Weinberg 1961), was 
the realization and the exploitation of the link between the generation 
of the one-parameter symmetry group (3 X by the algebraic derivation 
on A: 

(d/d\)(^\A))\ x=0 = (5 A) 

and the integral of the charge density jo of the conserved Noether cur- 
rent jp, d^jp = 0. 

Formally, such a link reads 

(5A) = lim ([Q R (t), A]} = lim ([Q R (0), A}}, (0.1) 

Q R (t) EE / rf 3 XJ (x,t). 
J\x\<R 

The above eq. (0.1), which is at the basis of the GSW proof for rela- 
tivistic systems (to be discussed below), is also the crucial ingredient 
of the Goldstone theorem for non-relativistic systems. In the latter 
case, a proper formulation and discussion of eq. (0.1), which shall be 
later referred to as the property of local generation of f3 x by the 
conserved Noether current j^, requires a special care, mainly be- 
cause the dynamics is not strictly local as in the relativistic case (see 
below). The point is that the check of the generation of the symme- 
try by the current density at equal times (which requires only the use 
of the canonical commutation relations) is not enough for establishing 
eq. (0.1). The statement (which may be found in the literature) that 
the independence of time of the commutator [Q(t), A] is guaranteed by 
the invariance of the Hamiltonian, since Q = i[H,Q] = 0, is not cor- 
rect. For the local generation at unequal times, the local property of 
the dynamics plays a crucial role. In fact, if the dynamics is strictly 
local, i.e. the time evolution of an operator localized in a compact 
set is still localized in a (possibly larger) compact set, the generation 
at equal times implies eq. (0.1). However, such an implication fails if 
the dynamics induces a long range derealization (see below) (Morchio 
and Strocchi 1985, Strocchi 2008) and the Goldstone theorem does not 
apply. 



7 



The first step in the discussion of eq. (0.1) is the condition of integra- 
bility of the charge density commutators. Technically, one must have 
that J(x, t) = i([,7o(x, t), A]) is absolutely integrable in x as a tem- 
pered distribution in t, i.e. after smearing with any test function h(t) 
of fast decrease (charge integrability condition). This condition is 
usually overlooked in the standard treatments of the non-relativistic 
Goldstone theorem, but it is needed for the continuity of the energy 
momentum spectrum in the neighborhood of k = 0, which is necessary 
for drawing the conclusions of the theorem (Swieca 1967, Morchio and 
Strocchi 1985, 1987, Strocchi 2008). 

The charge integrability condition is satisfied if the dynamics is 
sufficiently local, so that the (unequal time) commutators decay suffi- 
ciently fast in the limit of infinite space separations; the condition is 
obviously satisfied in the case of a strictly local dynamics, as it occurs 
in relativistic theories described by fields satisfying micro-causality. 

The localization of the dynamics plays a crucial role also for assuring 
the time independence of lim^_ >00 ([ Quit), A]), a necessary condition 
for its equality to the expectation of the time independent derivation 
SA. (anti) commutation relations are local, the issue is the effect of 
usually induced by the time evolution. 

Since, the continuity equation gives 

(d/dt)([j (x,t), A]) =div([j(x,t),A]>, 

the time independence of lim j fj^ 00 ([ Qn(t), A]) is guaranteed if the time 
evolution is sufficiently local, so that the Swieca condition (Swieca 
1967) is satisfied (s denotes the space dimension) 

lim |a| s - 1 ([j(x + a,0), a t (A)]} = 0. 

|a|— >oo 

Such a condition and more generally the needed time independence of 
lim^_ 5>00 ([ Qn(t), A]), which may actually require a weaker form of the 
Swieca condition (Strocchi 2005), is violated in the case of interactions 
described by a two body potential decaying as 1/r and this is the expla- 
nation of the evasion of the Goldstone theorem by Coulomb systems (as 
well as in the closely related Higgs mechanism in the Coulomb gauge). 



8 



In fact, roughly the logic of the theorem is the following: the time 
independence of 

J dxJ(x,t) = lim J due™* J(k,u), J(x,t) = i{[j (x,t), A]), 

implies that limk_>o j(k, u;) = XS(uj), (A a suitable constant). 

If there is symmetry breaking and f3 x is locally generated by j M , 
one has J dxJ(x,t) = (5 A) ^ 0, which implies A ^ 0. Thus, there 
cannot be a positive constant //, such that the energy spectrum of the 
intermediate states, which saturate the two point function J(x, t), is 
supported in {u;(k) > /i > 0}, in the neighborhood of k = 0; hence, 
there cannot be an energy gap /i. 

As a consequence of the above discussion one has the following 
(Lange 1966; Morchio and Strocchi 1987) 

Theorem 0.2 fGoldstone theorem for non-relativistic systems j 

// 

i) P x , A G R is a one-parameter internal symmetry group, 

ii) f3 x is locally generated by the conserved current j M (which transforms 
covariantly under space and time translations), eq. (0.1), 

Hi) f3 x is broken in the representation n with translationally invariant 
ground state vector i.e. there is A G A such that (5 A) ^ ; 
then there are quasi particle excitations with infinite lifetime in the 
limit k — > 0, i.e. particle-like excitations with energy width T(k) — > 
as k — > 0, and with energy u(k) — > as k — > ('Goldstone quasi 
particles^; the corresponding states have non-trivial projections in the 
subspaces {n(a t (A))^ , t G R} ; {ir(Q R )^ , R G R}. 

4. Goldstone theorem for relativistic quantum systems 

For relativistic quantum systems described by a local field algebra 
J 7 , the relevant representations are selected by the following physical 
conditions which strengthen the conditions I-IV discussed above, by 
incorporating relativistic invariance and relativistic locality: 
I. (Poincare covariance) The automorphisms a(a, A), (a G R 4 , A de- 
noting the elements of the restricted Lorentz group), which describe the 
transformations of the Poincare group are implemented by a strongly 
continuous group of unitary operators U(a,A(M)), M G SL(2, C), 



9 



II. (Relativistic spectral condition) H > 0, H 2 — P 2 > 0, 

III. (Vacuum state) There is a unique space-time translationally in- 
variant state \l/o (vacuum state) cyclic for the field algebra J 7 

IV. (Locality) The algebra J 7 satisfies relativistic locality, i.e. fields 
commute or anticommute at relatively spacelike points. 

In the case of relativistic local fields, the inevitable ultraviolet sin- 
gularities require a more careful regularization of the integral of the 
charge density j , e.g. 

Qr = 3o(fR, h) = J d 4 x f R (x) h(x ) j (x, x ), 

where / fl (x) = /(|x|/i2), / G P(R), f(y) = 1, if \y\ < 1, f(y) = 0, if 
\y\> 1 + e, he P(R), supp/i C [-5, 5], /i(0) = / dx h(x ) = 1. 

In the case of relativistic local fields the problems discussed above 
for non-relativistic systems, like the charge integrability condition, the 
local generation of /3 X and the time independence of the charge den- 
sity commutators do no longer arise thanks to locality. Then, one has 
(Goldstone, Salam and Weinberg 1962; Kastler,Robisson and Swieca 
1966) 

Theorem 0.3 (Goldstone theorem for relativistic local fields^ 

Let f3 x , X G R, be a one parameter group of internal symmetries locally 
generated by the covariant conserved current j^, i.e. \/F G J 7 

{6F)= lim i([Qr, F ]}; 

then, the symmetry breaking condition (SF) ^ implies that the Fourier 
transform J^(k) of the two point function (j^(x)F) contains a 5(k 2 ) 
contribution, i.e. there are Goldstone massless modes. 

The proof of the theorem is particularly simple if F is a scalar 
elementary field (p(x), i.e. a field transforming as a pointlike operator 
under the Poincare group: 

U(a, A) <p(x) U(a, A) -1 = ip(Ax + a). 

This was indeed the case considered in the original proof by Goldstone, 
Salam and Weinberg. In fact, the Poincare covariance of the two point 
function J^(x,y) = (j^x) (p(y)) implies that it has the following form 

Jp{x,y) = J„(x-y) =d fl J(x-y), 



10 



with J(x) a Lorentz invariant function. Then, the current conservation 
gives 

= d%(ar) = DJ(x). 

This means that J^(k) = Xk^5(k 2 ), with A ^ as a consequence of the 
symmetry breaking condition (5ip) ^ 0; hence, the intermediate states 
which saturate the two point function J^{x) have energy- momentum 
spectrum supported in k 2 = 0, i.e. they describe massless modes. 

The role of Lorentz covariance in the original proof (Goldstone, 
Salam and Weinberg 1962) has led to the belief that its failure provides 
the relevant mechanism for evading the conclusions of the theorem. 
This was indeed emphasized by Higgs (Higgs 1964) as the property of 
the Coulomb gauge which allowed the evasion of the Goldstone theo- 
rem for the Higgs mechanism. As stressed in the non-relativistic case, 
the crucial property is rather the localization property of the dynamics. 
As a matter of fact, the more general proof of the theorem, given by 
Kastler, Robinson and Swieca (Kastler, Robinson, Swieca 1967), covers 
the case of a symmetry breaking order parameter F without a pointlike 
structure, like a compound field or a polynomial of elementary fields, 
so that the two point function (j^(x) F) does not have simple transfor- 
mation properties under Lorentz boosts. 

In view of the crucial role of locality, a few comments may be rele- 
vant about its validity in the case of relativistic systems. Clearly, the 
algebra of observable fields and in particular the algebra A Q b s of their 
bounded functions, usually called the algebra of observables, must sat- 
isfy locality, since measurements of operators localized in spacelike sep- 
arated regions must be independent in the quantum mechanical sense. 

However, this does not imply that the states of the system, opera- 
tionally defined by their expectations of A Q b s , are local states, i.e. can 
be obtained by applying local fields to a vacuum state. Thus, one may 
have to use a non-local field algebra to reach such non-local states. In 
fact, this is the case of the charged states in quantum electrodynamics, 
whose description requires the non-local charged fields of the Coulomb 
gauge. Then, if the symmetry breaking is realized by the vacuum ex- 
pectation of a non-local field, the same problems of the non-relativistic 
systems arise and the same mechanism of evasion of the Goldstone 
theorem may take place. 



11 



5. Appendix 

In the seminal paper of 1964, Haag and Kastler emphasized the role 
of locality and, on the basis of mathematical and physical considera- 
tions, argued that the convenient mathematical setting for the descrip- 
tion of infinitely extended systems is that of C* -algebras; such algebras 
are (isomorphic to) norm closed algebras of operators in a Hilbert space, 
stable under the adjoing operation (Gelfand and Naimark 1943). 

The abstract definition of a C*-algebra A is that it is a linear asso- 
ciative algebra over the field C of complex numbers, with an involution 
* and equipped with a norm | | satisfying 

i) \\AB\\ < \\A\\ \\B\\ 

ii) \\A* A\\ = \\A\\* 

iii) A is complete with respect to the norm. 

In the following, we shall always consider unital C*-algebras A, i.e. 
with an identity 1, (1A = A = Al, \/A G A). 

A representation n of a C*-algebra A in a Hilbert space is a ho- 
momorphism of A into the C*-algebra £>(%„-) of the bounded operators 
in T-L n) namely a mapping which preserves all the algebraic relations, 
including the * (i.e. ir is linear, multiplicative n(AB) = tt(A) ir(B) and 
tt(A)* = ir(A*)). The property of preserving the * is spelled out by 
speaking of *-homomorphism, but for simplicity we shall omit the *. 
A homomorphism of A into the C*-algebra B which is one-to-one and 
onto (bijective) is an isomorphism] an isomorphism of A onto itself is 
called an automorphism. A vector \& G 'H- K is cyclic if n(A) ^ is dense 
in 'Hk- A representation n is faithful if kern = {0}, i.e. n(A) = 
implies A = 0. 

A state u> on A is a positive linear functional on A, i.e. MA, B G A 

u(XA + fiB) = Xu(A) + fj,u{B), A,/iGC, u{A*A)>0. 

A state won^l defines a representation n^A) of A in a Hilbert space 
"H w , i.e. a homomorphism of the C*-algebra A into the C* algebra of 
bounded operators in "H w . Moreover, contains a cyclic vector ^ 
such that cj(A) = (^ w , n u (A) ^E^). The representation is called the 
Gelfand-Naimark-Segal (GNS) representation defined by u and it is 
unique up to isomorphisms (Gelfand and Naimark 1943, Segal 1947). 

A state bj on A is called pure if it cannot be written as a convex linear 
combination of other states, and mixed otherwise. A mixed state ui on 



12 



A cannot be described by a state vector in an irreducible representation 
of A; in fact, it is described by a density matrix: 

u(A) = Ti( PuJ n(A)), MA e A, 

i i 

where Pi are one-dimensional orthogonal projections. 

Clearly, any vector $ G ^, with % the carrier of a representation 
7r of the C*-algebra A, defines a state on A, uiy(A) = (^, 7r(A) \I>) 
and the corresponding GNS representation is unitarily equivalent to ir. 

For quantum systems with a finite number 2iV of degrees of freedom, 
described by the canonical variables qk,Pk, k = 1,...N, the standard 
C*-algebra is the canonical Weyl algebra Aw, generated by the unitary 
operators U(a), V(/3), ct,/3 e R^, formally given by the exponentials 
U(a) = e l ^ kak<ik , V(f3) = e*^ fe/3fePfe , with the commutation relations in- 
duced by the canonical commutation relations of the canonical variables 
qk,Pk- in particular, U(a), V(j3) define TV-parameter abelian unitary 
groups. A representation n of the Weyl algebra is regular if n(U(a)), 
7i(V(f3)) are weakly continuous in a, (3, i.e. their matrix elements are 
continuous functions of a, (3; regularity is a necessary and sufficient 
condition for the existence of the generators, i.e. of the canonical vari- 
ables qk,Pk (as unbounded operators in / H n ). The Stone-von Neumann 
theorem (Stone 1930, von Neumann 1931) states that, up to unitary 
equivalence, there is only one regular irreducible representation of the 
canonical Weyl algebra. This means that there is only one (regular) 
realization of a system described by a canonical Weyl algebra. This im- 
plies that, for finite dimensional quantum systems, symmetries defined 
on the canonical variables q,p can never be broken (Strocchi 2008); the 
standard argument that symmetry breaking is prevented by tunneling 
(Stanford Encyclopedia of Philosophy 2008) looks less general and in 
fact does not apply to finite dimensional spin systems, for which an 
analog of the Stone-von Neumann uniqueness theorem may be proved. 

As advocated by Haag and Kastler, infinitely extended systems 
should be described by local C*-algebras, i.e. by norm closed *-algebras 
generated by operators which are localized. 

In the non-relativistic case, one may consider the canonical field 
operators ipix), V ; *( x ) ) localized at x, smear them with test functions 



13 



/, g G (^(R 3 ), of compact support (i.e. G £>(R 3 )) or of fast decrease 
(i.e. G 5(R 3 )), and take the C*-algebra A generated by the exponen- 
tials 

U(f) = e iWf)+rU) \ V(g) = e^ 9) ~ r{9) . 

This algebra represents the infinite dimensional analog of the Weyl 
algebra; the Stone-von Neumann theorem does not apply and in fact 
there are many inequivalent representations. 

In the case of relativistic systems, Haag and Kastler proposed to 
define a quantum field theory by the association of a C*-algebra A(0) 
to each bounded region O, typically a double cone, in Minkowski space, 
with the following setting: i) isotony: if D 2 , then A(Oi) C A(0<z); 
ii) locality: if 0\ and 2 are completely spacelike with respect to each 
other, then A(Oi) and .4.(02) commute; iii) the norm completion A of 
the set theoretical union of all A(0), called the algebra of observ- 
ables, contains all the observables; iv) relativistic covariance: the inho- 
mogeneous Lorentz group is represented by automorphisms A — > oll{A) 
such that a L (A(0)) = A(O l ), VA G A, with L the image of O under 
the inhomogeneous Lorentz transformation L. 

The algebra A(O) has the meaning of the observables which can 
be measured in the space-time region O. The C* structure emphasizes 
that fact that, from an operational point of view, observables should 
be described by bounded operator (Segal 1947, Strocchi 2005). From a 
mathematical point of view, there is a big advantage in using bounded 
operators, since e.g. domain questions do not arise. For simplicity, we 
shall not spell out the distinction between the union of the A(0), called 
the local observable algebra, and its completion called the quasilocal 
algebra of observables. Furthermore, it is taken for granted that A 
admits a faithful algebraically irreducible representation. 

The standard formulation of quantum field theory makes use of local 
quantum fields, which are unbounded operators and therefore cannot 
be elements of a C*-algebra. Furthermore, there are quantum fields 
which do not describe observable quantities; e.g. a fermion field is not 
an observable field, whereas so is the electromagnetic field F liU . Thus, 
the algebra T generated by the local fields contains the subalgebra T bs 
of observables fields; the C*-algebra of observables should be obtained 
by considering bounded functions of the observable fields. For example, 
e tF v(f\ with / G £>(R 4 ), is a bounded operator and may be considered 



14 



as an element of the C*-algebra of observables. 

In the Haag-Kastler approach, called the algebraic approach to quan- 
tum field theory, a symmetry is an automorphism of the algebra of ob- 
servables and it is broken in a representation ir if it is not implementable 
by a unitary operator there. 

Acknowledgments. I am indebted to Riccardo Guida for useful com- 
ments in the preparation of this note. 



15 



REFERENCES 
RW. Anderson, Phys. Rev. 130, 439 (1963) 

T. Eguchi and K. Nishijima, Broken Symmetry. Selected papers by Y. 
Nambu, World Scientific 1995 

I. Gelfand and M.A. Naimark, On the imbedding of normed rings into 
the ring of operators in a Hilbert space, Mat. Shorn., N.S., 12, 197-217 
(1943) 

J. Goldstone, Field Theories with Superconductor Solutions, Nuovo 
Cim. 19, 154-164 (1961) 

J. Goldstone, A. Salam and S. Weinberg, Broken Symmetries, Phys. 
Rev. 127, 965-970 (1962) 

R. Haag, Local Quantum Physics, Springer 1996 

R. Haag and D. Kastler, Algebraic approach to quantum field theory, 
Jour. Math. Phys. 5, 848-861 (1964) 

W. Heisenberg, Zut Theorie des Ferromagnetism, Z. Physik 49, 619-636 
(1928) 

P.W. Higgs, Broken Symmetries, Massless Particles and Gauge Fields, 
Phys. Lett. 12, 132-133 (1964) 

D. Kastler, D.W. Robinson and J. A. Swieca, Conserved currents and 
associated symmetries; Goldstone's theorem, Comm. Math. Phys. 2, 
108-120 (1962) 

R.V. Lange, Nonrelativistic Theorem Analogous to the Goldstone The- 
orem, Phys. Rev. 146, 301-303 (1966) 

G. Morchio and F. Strocchi, Infrared problem, Higgs phenomenon and 
long range interactions, Lectures at the Erice School of Mathematical 
Physics, Erice 1985, Fundamental Problems of Gauge Field Theories, 
G. Velo and A.S. Wightman eds. Plenum 1986 

G. Morchio and F. Strocchi, Mathematical structures for long-range 
dynamics and symmetry breaking, Jour. Math. Phys. 28, 622-735 
(1987) 

Y. Nambu, Axial vector current conservation in weak interactions, 
Phys. Rev. Lett. 4. 380-382 (1960) 

Y. Nambu, Quasiparticles and Gauge Invariance in the Theory of Su- 
perconductivity, Phys. Rev. 117, 648-663 (1960) 

I. Segal, Postulates of general quantum mechanics, Annals of Math., 
48, 930-948 (1947) 



16 



Stanford Encyclopedia of Philosophy, Symmetry and Symmetry Break- 
ing, 2008 

M.H. Stone, Linear transformations in Hilbert space, III. Operational 
methods and group theory, Proc. Nat. Acad. Sci. U.S.A., 16, 172-175 
(1930) 

F. Strocchi, Symmetry Breaking, Springer, 2005, 2nd ed. 2008 

F. Strocchi, An Introduction to the Mathematical Structure of Quantum 

Mechanics, World Scientific 2005, 2nd enlarged edition 2010 

J. A. Swieca, Range of forces and broken symmetries in many-body 

systems, Comm. Math. Phys. 4, 1-7 (1967) 

J. von Neumann, Die Eindeutigkeit der Schroedingerschen Operatoren, 
Math Ann. 104, 570-578 (1931) 

E.P. Wigner, Group Theory and Its Applications to the Quantum Me- 
chanics of Atomic Spectra, Academic Press 1959